You are on page 1of 8

Anaerobic Migrating Blanket Reactor

Treatment of Low-Strength Wastewater at


Low Temperatures
Largus T. Angenent, Gouranga C. Banik, Shihwu Sung

ABSTRACT: The feasibility of the compartmentalized anaerobic meet effluent quality standards, such as for effluent nitrogen and
migrating blanket reactor (AMBR) was studied for the treatment of low- phosphorus levels (Rodrigues et al., 2001).
strength soluble wastewater under low-temperature conditions. During For full-scale treatment of low-strength wastewater, the upflow
an operating period of 186 days, a 20-L AMBR was fed nonfat dry milk anaerobic sludge blanket (UASB) reactor was used widely
substrate as a synthetic wastewater at low temperatures (15 and 20 C).
(Hulshoff Pol et al., 1997, and Lettinga et al., 1993). The UASB
The concentration of the influent was constant at chemical oxygen
demand (COD) and 5-day biological oxygen demand (BOD5) reactor consists of a single vessel with a hydraulic upflow pattern,
concentrations of 600 and 285 mg/L, respectively. The soluble COD and hence a gassolids separator system and a feed-distribution
(SCOD) removal efficiency was 73% at the end of the operating period system are required to retain biomass and to distribute influent
(15 C) at a 4-hour hydraulic retention time (HRT), while the total COD evenly, respectively. Kato (1994) showed in laboratory-scale stud-
(TCOD) removal efficiency was 59%. At a 4-hour HRT, staged ies that the expanded granular sludge bed (EGSB) reactor, which
conditions promoted complete removal of propionic acid in the final resembles a UASB reactor with a higher upflow velocity, was
compartments of the reactor. The specific methanogenic activity of more efficient compared with the UASB reactor for this type of
granules increased slowly until the end of the operating time, improving wastewater. Evidently, a higher mixing intensity in EGSB reactors
the removal rate. Biomass was retained effectively, as evidenced by the
because of a higher upflow velocity decreased transport diffusion
solids retention time (SRT) that was always greater than 50 days even
during step decreases of the reactor HRT from 12 hours to 4 hours. A
limitations of substrate into the granules. Feasibility of this tech-
long SRT also promoted system stability during changes in flow, which nology was confirmed with pilot-scale studies treating low-
was observed by SCOD removal efficiencies staying greater than 70%. strength malting wastewater at 13 C and 20 C (Rebac et al.,
During a hydraulic stress test, the HRT was reduced from 4 hours to 1 1997).
hour for one day (24 HRTs) in which volatile suspended solids (VSS) in Mechanically mixed, anaerobic sequencing batch reactors (AS-
the effluent increased from an average background level of 8.7 g/d to 35 BRs) were also thought to be an attractive option for the treatment
g/d and the SRT decreased from 50.5 days to 12.6 days. However, of low-strength, low-temperature wastewater, because increased
mixed liquor volatile suspended solids concentration decreased only by mixing conditions are combined with ideal conditions for biomass
1 g/L, and hence a similar COD removal efficiency and biogas production
settling and retention in the reactor. In addition, because of the
was found one day after the hydraulic stress (as compared to one day
before the hydraulic stress). Water Environ. Res., 73, 567 (2001).
absence of a hydraulic upflow pattern, gassolids separator and
feed-distribution systems are superfluous, which simplified the
KEYWORDS: anaerobic treatment, low-strength wastewater, low-tem- reactor configuration compared with UASB and EGSB systems.
perature conditions, compartmentalized reactor, granules, staged treatment, Despite eliminating a hydraulic upflow pattern, the ASBR was able
anaerobic migrating blanket reactor, methanogenesis.
to develop and grow granular biomass (Sung and Dague, 1995, and
Wirtz and Dague, 1996), which protects the strict anaerobes from
Introduction oxygen toxicity in low-strength, low-temperature wastewater treat-
Advantages of anaerobic pretreatment over aerobic treatment ment (Kato, 1994). Studies by Dague et al. (1998) on ASBR
for low-strength wastewater include decreased sludge production treatment of nonfat dry milk wastewater at temperatures ranging
and lower energy requirements, which result in decreased operat- from 5 to 25 C showed that solids retention times (SRTs) were
ing costs (Mergaert et al., 1992). Low-strength wastewater, such as always exceeding 25 days and hence could offset the low growth
domestic and food-processing wastewater, is often discharged at rates of biomass at these low temperatures. At a temperature of 5
ambient temperatures (15 to 20 C) and heating the wastewater to C and a 6-hour hydraulic retention time (HRT), the ASBR still
maintain mesophilic conditions (35 C) for anaerobic treatment achieved a total chemical oxygen demand (TCOD) removal effi-
introduces large energy requirements. This would eliminate sav- ciency of 60%. An advantage of the ASBR technology was the
ings in operating cost for the anaerobic system completely. There- alternating higher and lower substrate levels (feast and famine
fore, pretreatment of low-strength wastewater is more attractive conditions) during the operating cycle, which resulted in increased
under low-temperature conditions. Recently, several laboratory- biomass settling and retention before effluent decanting and in-
and pilot-scale studies revealed promising results for high-rate creased substrate utilization rates just after feeding (Dague et al.,
anaerobic treatment of low-strength wastewater at temperatures as 1998). Increased substrate utilization is preferred, as for low-
low as 3 to 5 C (Dague et al., 1998; Lettinga et al., 1999; and strength wastewater substrate utilization rates are generally low,
Rebac et al., 1997). Most often an aerobic polishing (posttreat- dictated by Monod kinetics (Monod, 1949).
ment) step is needed after the anaerobic system (pretreatment) to The kinetic advantage of relatively high substrate levels can also

September/October 2001 567


Angenent et al.

be obtained with a continuously fed staged reactor configuration.


Lettinga et al. (1999), for example, used two EGSB reactors in
series to gain from higher substrate concentrations in the first
EGSB reactor, when feeding a low-strength volatile fatty acid
(VFA) synthetic wastewater at a 1.5-hour HRT and a temperature
of 8 C. These researchers found also that staging was beneficial
for the following reasons: (1) more favorable conditions for pro-
pionic acid degradation in the staged-reactor configuration because
of low acetic acid and hydrogen levels in the second stage; and (2)
relative high levels of bacteria in the first stage of the system Figure 1Schematic diagram of AMBR (EBS effluent
(while methanogens were still present) and relative high levels of baffle system).
methanogens in the second stage (biomass staging), which im-
proved biodegradation kinetics due to each major group of micro-
organisms existing in an optimal environment (Lettinga et al., Before the flow was reversed, the second compartment was step
1999, and van Lier et al., 1997). fed for 4 hours (after termination of feeding the initial compart-
Another continuously fed staged reactor, the anaerobic migrat- ment) to prevent a breakthrough of substrate. Sufficient contact
ing blanket reactor (AMBR) was developed, because there is still between substrate and biomass was maintained using intermittent,
a need for simple and cost-effective high-rate anaerobic treatment gentle mixing. Mixers (model 5vb, EMI Inc., Clinton, Connecti-
systems for small and medium sized communities and industries, cut) were mounted with paddles to ensure gentle mixing. At a
especially for treating low-strength wastewater (Hulshoff Pol et rotational speed of 60 r/min, these paddles produced a root mean
al., 1997). The AMBR utilizes the advantages of the ASBR, such square velocity gradient (G) of 77 s1 in a 5-L compartment, as
as mechanical mixing, biomass retention, a simple design (no determined by a rotating torque meter (Bex-O-Meter, model 38,
gassolids separation and feed-distribution systems required be- The Bex Company, San Fransisco, California) described in Sajjad
cause of the absence of a hydraulic upflow pattern), and granula- and Cleasby (1995). The compartments were mixed equally for 10
tion (Angenent and Sung, 2001). In addition, the HRT in a con- seconds every 4 minutes at 60 r/min for the first 155 days of
tinuously fed AMBR can be shortened (and thus the reactor operation. At day 156, the mixing frequency was doubled to once
volume can be decreased) compared with the batch-fed ASBR. every 2 minutes in the three initial compartments, and the final
Relatively long HRTs due to physical limitations of the system (a compartment was mixed every 4 minutes to prevent excessive
combination of a batch-fed operation and a relatively tall granular biomass loss. Increasing the mixing frequency of the initial com-
blanket) are required for the ASBR, because a reduction of the partments was implemented to decrease transport limitation and
6-hour HRT for ASBRs treating low-strength wastewater was not short-circuiting of substrate at higher organic loading rates. All
anticipated (Dague et al., 1998). pumps used were Masterflex pumps (Cole Parmer Instrument Co.,
Because of the staged and continuously fed configuration of the Chicago, Illinois).
AMBR (and thus a potential for short HRTs) in combination with The gas collection systems consisted of an observation bottle, a
a simple design, the feasibility of the AMBR system for the gas sampling port, and a wet-test gas meter (GCA, Precision
treatment of highly soluble, low-strength wastewater was studied. Scientific, Chicago, Illinois). The biogas was directly discharged
The AMBR treatment was evaluated under low-temperature con- from the reactor to the gas collection system. A water head was
ditions (15 C) by monitoring reactor performance over a period of installed on the effluent tubes to prevent biogas from escaping
6 months, during which the HRT was reduced from 12 hours to 4 directly through the effluent ports. Timers (ChronTrol Corpora-
hours in a stepwise manner. Biomass retention, granular size, and tion, San Diego, California) regulated the operation. An effluent
methanogenic activity of granules were also monitored during this baffle system was placed in front of the effluent ports to prevent
period. In addition, a sudden drop of HRT from 4 hours to 1 hour floating granules from leaving with the effluent (Figure 1).
was introduced to study the behavior of the AMBR under hydrau- Inoculum. Granular inoculum was obtained from laboratory-
lic-stress conditions, which are common for low-strength waste- scale ASBRs fed with the same substrate (Dague et al., 1998) and
water. was stored for 3 months at 5 1 C and for 1 month at 20 1 C
before operation. These granules were acclimated to psychrophilic
Materials and Methods temperatures as low as 5 1 C. At start-up, the mixed liquor
Anaerobic Migrating Blanket Reactor. The AMBR, with an volatile suspended solids (MLVSS), which is an indication of the
active volume of 20 L (inside width: 140 mm; inside length: 580 amount of biomass in the reactor, was 20 g/L.
mm), was divided into four compartments (Figure 1). Baffles were Substrate. A concentrated substrate stock solution, consisting
placed between the compartments to reduce short-circuiting of of nonfat dry milk (962 mg/g chemical oxygen demand [COD]),
substrate. The space between a baffle and the inside wall was 10 sodium bicarbonate (962 mg/g feed COD) and trace elements
mm to prevent clogging by large granules. The temperature of the (ferrous chloride, FeCl24H2O: 17.1 mg/g COD, nickel chloride,
AMBR was kept constant at 20 1 C for the first 87 days of NiCl26H2O: 1.9 mg/g COD, cobaltous chloride, CoCl26H2O: 1.9
operation. At day 88, the temperature was decreased to 15 1 C, mg/g COD, manganese chloride, MnCl24H2O: 1.7 mg/g COD,
and precooling of influent became necessary. The flow over the and zinc chloride, ZnCl2: 1.3 mg/g COD), was stored at 4 1 C
horizontal plane of the reactor was reversed once a day. A daily and was mixed during feeding to obtain a constant loading rate.
change in flow direction was chosen to prevent a pH drop due to Makeup water was added to the concentrated substrate just before
VFA buildup in the initial compartment and to prevent unequal introduction to the reactor to obtain a COD and 5-day biochemical
biomass levels due to anticipated biomass migration between oxygen demand (BOD5) concentration of 600 mg/L and 285 mg/L,
compartments. respectively (the BOD5/COD ratio was 0.49). The soluble fraction

568 Water Environment Research, Volume 73, Number 5


Angenent et al.

of COD (SCOD) of the nonfat dry milk substrate solution was on obtained after dividing the biogas production with the wet volume
average 8.4% smaller than the TCOD fraction. The sulfate con- of the reactor and converting it to the methane percentage that was
centration of the influent (mainly from the city of Ames, Iowa, tap present in the biogas. Therefore, the SMPR was expressed as
water) was approximately 110 mg/L (COD/sulfate ratio was 5.5). volume of methane production per reactor volume per day (L/Ld).
Analyses. Concentrations of nitrogen, methane, and carbon The methane that left the reactor in a soluble form with the effluent
dioxide in the headspaces of the AMBR and activity bottles were was then calculated. First, the solubility of methane in water at 15
measured using a gas chromatograph (model 350, Gow-Mac In- or 20 C was estimated using Henrys law (Perry et al., 1997).
struments, Co., Bridgewater, New Jersey) with a thermal conduc- Second, the methane quantity per day that left the reactor through
tivity detector (column: 1.7 m x 3 mm stainless steel Poropack Q its effluent was calculated and was corrected to STP and converted
0.177/0.149 mm [80/100 mesh]; carrier gas: helium). The individ- to SMPR. Theoretically, 0.35 L of methane has 1 g of COD, which
ual VFA levels were measured with an ion chromatograph (model also means 1 g of COD can potentially generate 0.35 L of methane
DX-500, Dionex, Sunnyvale, California) containing a CD 20 con- at STP (when biomass growth is ignored). Hence, the methane-
ductivity detector and an anion micromembrane suppressor (col- based COD (MCOD) removal efficiency can be calculated by the
umn: Ion Pac ICE-As1; eluant: 0.8 to 1.0 mM heptafluorobutyric total SMPR (methane that left the reactor through the gas meter
acid). For VFA analyses, samples were first acidified with hydro- and the effluent) using the following formula:
chloric acid. The total alkalinity, total VFAs, TCOD, SCOD, total
suspended solids (TSS), volatile suspended solids (VSS), and SMPR
MCOD removal efficiency, % (100)
sludge volume index tests were performed according to procedures (VLR)(0.35)
described in Standard Methods (APHA et al., 1995). Effluent
samples were taken at the midpoint of the time interval between Results and Discussion
reversals of flow. Operating Conditions. At the start of the operating period, the
Biomass Characteristics. The specific methanogenic activity of reactor was fed at a 12-hour HRT and a VLR of 1.25 g COD/Ld.
biomass was assessed at 35 1 C using the headspace method These loading rates at a temperature of 20 C were possible,
according to tests described by Rinzema et al. (1988). To analyze because the inoculum was already acclimated to similar environ-
the change in size of the granules during the operational time, the mental conditions. The concentration of the influent was kept
arithmetic mean diameter [Sum (d/n), where d is the granule constant at 600 mg COD/L (as TCOD) during the entire operating
diameter and n is the number of granules] and area-weighted mean period and decreasing the HRT resulted in an increase in the VLR.
diameter [Sum(d3)/Sum(d2)] were calculated by automated image At the end of operating time, the HRT was 4 hours, which resulted
analysis. Mixed liquor samples were collected six times through- in a VLR of 3.5 g COD/Ld (Figure 2a). The operating length of
out the course of the study and diluted to obtain fully suspended each different HRT was chosen based on TCOD, SCOD, and
granular biomass samples for image analysis. Next, 1.75 mL of MCOD removal performance. If two or more data points were
this sample was added to a specially prepared slide consisting of within 5% margin after the system was operated for at least 10
two, 3-mm thick glass sheets cemented together, with a 25-mm HRTs, a quasi steady state was determined. True steady-state
diameter hole in the top sheet. The image analysis setup contained conditions were not achieved at the HRTs of 12, 8, and 4 hours,
a black and white video camera (series 68, Dage-MTI, Michigan because the MLVSS and SRT were still increasing slowly while
City, Indiana), a microscope (model SZH, Olympus, Melville, the HRT was decreased again (Figure 2c). It needs to be empha-
New York), and a personal computer with Quartz PCI Imaging sized that the AMBR is a dynamic process in which substrate
software (Quartz Imaging Corporation, Vancouver, British Colum- concentrations and environmental conditions change over time
bia, Canada). Some manual editing of the image was necessary to because of reversing the flow.
separate adjacent granules. Particles smaller than 0.1 mm were not Figure 2 illustrates the operating conditions during the operating
included in the calculations of the size distribution (Grotenhuis et period. Biomass levels in the reactor stayed at approximately 23
al., 1991). g/L during this period (Figure 2b). Meanwhile, VSS losses with
The biomass migration rate was determined by the decline in the effluent oscillated around a mean of 4.4 g/d (standard error,
MLVSS from the initial compartment during the time of feeding. SE 1.9; number of datapoints, n 18) during the operating
The food/microorganism (F/M) ratio was calculated by dividing period (not including the hydraulic stress data). The oscillation
the mass of COD fed per day with the total VSS in the reactor. The pattern was the result of temporary solids loss in the effluent after
SRT was determined by dividing the total VSS in the reactor with a temperature decrease at day 88 and after the increased flows due
the daily loss of VSS in the effluent. to HRT changes. Even during the temporary biomass loss in-
Assessment of Reactor Performance. To obtain information creases, biomass retention was high and the SRT remained greater
on the performance of the system, COD removal efficiencies were than 50 days. After a period of acclimation, VSS levels in the
based on COD concentrations measured in the influent and effluent effluent decreased and the SRT increased to values greater than
and methane production. The volumetric loading rate (VLR) was 100 days. The HRT could then be decreased again (by increasing
calculated as the mass of COD fed to the system per reactor the flow), which resulted in temporarily lower SRTs of 50 days
volume per day (g COD/Ld). Data for the SCOD and TCOD (Figure 2c). Because the MLVSS levels remained constant, F:M
removal efficiencies were obtained by measuring the COD con- during the operating period increased from 0.05 to 0.18 g COD/g
centrations of the feed and the effluent. Effluent samples and VSSd, depending on VLR (Figure 2c). Meanwhile, pH levels in
samples from individual compartments were taken at the midpoint the initial compartment measured at the midpoint between rever-
of the time interval between two reversals of flow. Methane sals of flow (12 hours of feeding the initial compartment) were
production was determined as follows. First, the biogas production constant over the operating period with a mean of 6.7 (SE 0.1;
(gas meter) was corrected to standard temperature and pressure n 30). The alkalinity in the effluent remained constant as well
(STP). Next, the standard methane production rate (SMPR) was (mean 0.53 g/L as calcium carbonate; SE 0.04; n 18).

September/October 2001 569


Angenent et al.

Therefore, the increase in F:M showed no decrease in pH level in


the initial compartment, which indicates that stable conditions
prevailed in the reactor.
Anaerobic Migrating Blanket Reactor Performance. Figure
3 illustrates the reactor performance over the operating period. At
a 12-hour HRT, biogas production (measured with the gas meter)
reached a plateau within a month of feeding the reactor at a
temperature of 20 C (Figure 3a). At this period (days 24 to 43),
the AMBR achieved SCOD, TCOD, and MCOD removal efficien-
cies of 93.9% (SE 1.8; n 3; average effluent SCOD concen-
tration, eSCOD 34.6 mg/L), 81.3% (SE 5.0; n 3; average
effluent TCOD concentration, eTCOD 105.5 mg/L), and 61.0%
(SE 7.9; n 20), respectively (Figure 3b). The MCOD removal
efficiencies were lower than the TCOD removal efficiencies pri-
marily because of biomass accumulation in the reactor (Figure 2b).
The lower MCOD removal efficiency can also be partly explained

Figure 3Reactor performance: (a) biogas and methane


production, (b) COD removal efficiency, and (c) biomass
activity and granular size. Results during the hydraulic
stress test are indicated by peak flow.

by sulfide formation by sulfate-reducing bacteria and the subse-


quent sulfide loss, and thus COD loss, through stripping of hydro-
gen sulfide or precipitation of metal sulfides. A theoretical maxi-
mum of 10% of the SCOD removed can be attributed to this
phenomenon when 100 mg/L sulfate was reduced and all sulfide
was stripped or precipitated. An increased VLR resulting from the
HRT decrease from 12 hours to 8 hours increased the biogas
production slowly until day 78. Feeding was terminated for 2 days
from day 78, which resulted in a reduced biogas production. After
feeding the system again for 7 days, the biogas production reached
a level similar to that before the termination of feed.
Reducing the operating temperature from 20 C to 15 C on day
88, decreased the biogas production (as measured with the gas
meter) from approximately 8 L/d to 6 L/d and the SMPR (gas
Figure 2Operating conditions: (a) loading conditions, meter) from approximately 0.3 L/Ld to 0.2 L/Ld (Figure 3a). The
(b) biomass concentrations in reactor and effluent, and increased level of soluble methane in the effluent at 15 C com-
(c) SRT and F:M. Results during the hydraulic stress test pared with 20 C did not correct entirely for the decrease in
are indicated by peak flow. measured methane and biogas production. After decreasing the

570 Water Environment Research, Volume 73, Number 5


Angenent et al.

temperature, the SMPR for the soluble methane fraction accounted production can be reduced greatly, as the data suggest a true
for one third of the total SMPR at a temperature of 15 C (Figure growth yield of 0.16 g biomass COD formed/g SCOD removed
3a). The biogas production remained at a similar level for the rest over the final three data points at a 6-hour HRT (15 C). This is in
of the 8-hour HRT period (days 94 to 108) during which the agreement with ASBR data that showed a true growth yield of
SCOD, TCOD, and MCOD removal efficiencies were 82.1% 0.145 g COD/g COD at 15 C (Banik et al., 1998). In contrast,
(SE 1.0; n 3; eSCOD 96.3 mg/L), 70.3% (SE 3.0; n typical activated-sludge growth yields are greater than (0.6 g
3; eTCOD 159.8 mg/L), and 53.4% (SE 7.3; n 13), COD/g COD) (Grady et al., 1999). In contrast to the procedures
respectively. These removal efficiencies decreased further when used by Barbosa and Sant Anna (1989), the AMBR of the current
the HRT was reduced from 8 hours to 6 hours. At the end of the study was operated with a highly soluble synthetic wastewater. For
6-hour HRT period (days 157 to 165), the biogas production nonsettled domestic wastewater that contains relatively high levels
increased slowly to approximately 7 L/d (as measured with the gas of suspended solids, the AMBR can be operated at a longer HRT
meter). During this period, the AMBR achieved SCOD, TCOD, to facilitate primary solids separation and the possibility to main-
and MCOD removal efficiencies of 75.9% (n 2; eSCOD tain flocculent biomass instead of granular biomass to promote
136.1 mg/L), 62.1% (n 2; eTCOD 213.7 mg/L), and 42.8% suspended and colloidal solids destruction. For this type of waste-
(SE 2.1; n 9), respectively, which were lower than those at the water, hydrolysis and acidogenesis rather than methanogenesis
8-hour HRT and temperature of 15 C (Figure 3b). become the rate-limiting steps, and hence to provide for solids
The biogas production increased to approximately 8 L/d (gas destruction, the entrapment of suspended solids and long SRTs
meter) when the HRT was further decreased to 4 hours, indicating become important (Switzenbaum and Grady, 1986). An AMBR
that the system was increasing its removal rates. An increase in operated with (partly) flocculent biomass may provide for suffi-
removal rate was also determined by the specific methanogenic cient biomass and solids retention, because this study with granular
activity of the granules, which increased slowly for the final 100 biomass showed an SRT that was longer than 100 days during
days of operation (Figure 3c). The increase in removal rate resulted pseudo-steady-state conditions. However, further studies are
in only slightly lower COD removal efficiencies at a 4-hour HRT needed to elucidate the feasibility for AMBR treatment of non-
compared to the 6-hour HRT. The SCOD, TCOD, and MCOD settled domestic wastewater.
removal efficiencies for the two data points (days 172 and 184) at Besides reducing nutrient levels in the effluent, posttreatment of
a 4-hour HRT before the hydraulic stress test were 73.0% (n 2; anaerobic reactor effluent is needed to reduce the mean SCOD
eSCOD 163.9 mg/L), 59.2% (n 2; eTCOD 248.4 mg/L), concentrations of 106 mg/L (SE 50; n 19) as found in this
and 35.6% (SE 1.7; n 11), respectively. The lower SCOD study. The soluble organic material in the effluent consisted of
removal efficiencies were a result of an increase in the VFA VFAs with a mean total concentration of 39 mg/L as acetate (SE
concentrations of the effluent from 30 mg/L as acetic acid (n 2) 14; n 21). The remaining soluble organic material may have
to 60 mg/L (n 2) during the HRT decrease. partly consisted of soluble microbial products (Langenhoff and
The operating performance showed an SCOD removal effi- Stuckey, 2000). Most soluble microbial products are not, or
ciency greater than 70% throughout the operating period, while the slowly, biodegradable and, therefore, are not measured with BOD5
TCOD removal efficiency remained at approximately 60% for a tests. Hence, BOD5 measurements and removal efficiencies would
4-hour HRT. These values did not dramatically decrease after HRT have given an additional evaluation of the performance of the
changes, and hence the system showed stability after flow in- AMBR, especially as the nonfat dry milk substrate was not fully
creases. During each period of a 12, 8, and 6-hour HRT, the degraded under aerobic conditions within 5 days (BOD5/COD
performance of the AMBR improved slightly. Therefore, the per- ratio of 0.49).
formance at a 4-hour HRT was expected to have improved at a Biomass Characteristics. An increase in the size of the gran-
longer operating period, because the biogas production and spe- ules over time is illustrated in Figure 3c. At the end of the
cific methanogenic activity of the granules were increasing slowly operating period, the arithmetic and area-weighted mean diameter
at the end of the operating time. Dague et al. (1998) performed were 1.1 and 3.0 mm, respectively. Also, the area distribution of
similar studies with ASBRs fed with a similar nonfat dry milk the granules in Figure 4 shows an increase in granular diameter
substrate. At a 6-hour HRT and a temperature of 15 C, these over the operating time, during which the graph at day 185 had
workers found an SCOD removal efficiency of 81.5%. A slightly shifted to the right compared with the graph of the seed granules
lower SCOD removal efficiency of 75.9% at similar conditions and the graph at day 105. At day 185, 40% of the projected
was achieved during the current study. Although the ASBR granular area was due to particles between 3.16 and 5.62 mm in
achieved a higher SCOD removal efficiency, the HRT could not be diameter. For the seed sludge no particles with a diameter between
reduced from 6 hours to 4 hours because of physical limitations of 3.16 and 5.62 mm were found (Figure 4). A selection for larger
the batch-fed system. In contrast, the HRT of the AMBR was granules occurred because of growth and an increasing washing
reduced to 4 hours and was even decreased to 1 hour during the out of smaller biomass at higher flows. It needs to be realized that
hydraulic stress test. selection of larger granules occurred despite the absence of a
Pretreatment of raw domestic wastewater by a UASB system at hydraulic upflow pattern. At the end of the operation, the migration
ambient temperatures was studied by Barbosa and Sant Anna rate of large granules was virtually zero because of their good
(1989). Although solids were present in their wastewater, the settling characteristics (sludge volume index was 18.7 mL/g) and
BOD5 and COD concentrations were 357 and 627 mg/L, respec- a compartmentalized configuration of the AMBR. This absence of
tively, which were close to those of the current study (285 and 600 migration became ubiquitous because a difference in granular
mg/L, respectively). Similar to this previous study, current results structure was observed between the compartments. The surfaces of
suggest that the AMBR as a pretreatment step can be effective in granules from the outside compartments showed an abundance of
removing COD, and thus can reduce energy requirements and add white colonies (75 m in diameter), consisting of microorganisms
stability to an aerobic posttreatment facility. Moreover, sludge with a different morphology (microscopic observations of the

September/October 2001 571


Angenent et al.

colonies showed a homogeneous population), while the granules


from the inside compartments continued to have a smooth black
surface. Because of high concentrations of sulfate in the influent
(110 mg/L), these microcolonies, presumably, consisted of sulfate-
reducing bacteria as was also described by Sekiguchi et al. (1999),
who discovered colonies consisting of Desulfobulbus spp. on gran-
ules by using fluorescence in situ hybridization.
The specific methanogenic activity of biomass for the outside
compartments from the AMBR was 0.53 g COD/g VSSd (stan-
dard deviation, SD 0.01; number of measurements, m 2) and
was 0.58 g COD/g VSSd (SD 0.01; m 3) for the inside
compartments. Hence, at the end of the study a small difference in
specific methanogenic activity was found between the granules of
the inside and outside compartments (t-test: 95% significance
level), which resulted in biomass staging. This nomenclature
indicates that different microbial communities developed over the
length of a plug-flow reactor or in a sequence of compartments.
For example, biomass staging in a two-stage EGSB reactor fed
with partly acidified substrate at temperatures as low as 8 C,
resulted in a dominant acidogenic population in the first stage and
dominant acetogenic and methanogenic populations in the second
stage (van Lier et al., 1997). Results of the current study suggest,
however, a much smaller difference in biomass between the com-
partments. The reversing flow scheme is the most likely reason
why biomass staging was limited for the AMBR compared with
the unidirectionally fed, two-stage EGSB system. The above given
specific methanogenic activity was measured at 35 C, which
verified that mesophilic methanogens remain active over the op-
erating period despite low-temperature conditions, as was also
found by Lettinga et al. (1999), Banik et al. (1997), and Langen-
hoff and Stuckey (2000).
Hydraulic Stress. At day 185, the HRT was decreased from 4
hours to 1 hour to study the effect of a temporary increase in
wastewater flow (hydraulic stress). During this 24-hour period (24
HRTs), a total of 480 L of feed was applied to both flow directions
(240 L was fed to each direction for 12 hours). One day before the
hydraulic stress test, at day 184, staged environmental conditions
in the AMBR were apparent with relatively high concentrations of
acetic acid and SCOD in the initial compartment (compartment 1)
and low concentrations in the final compartment (Figure 5a). The Figure 5Volatile fatty acid and SCOD concentration at
appearance of formic acid and propionic acid in the initial com- the midpoint between intervals of flow per compartment:
partment indicated the occurrence of substrate staging. In this (a) t 184 days and general operating conditions at a
4-hour HRT; (b) t 185 days and hydraulic stress at a
1-hour HRT; and (c) t 186 days and day after hydraulic
stress at a 4-hour HRT.

paper, substrate staging indicates that different substrate levels


(and thus environmental conditions) are present over time or over
a spatial arrangement of the reactor configuration, regardless of
whether biomass staging occurs. Propionic acid concentrations that
were below the detection limit illustrated favorable propionic acid
degradation in compartments 2 to 4. Grobicki and Stuckey (1992),
van Lier et al. (1996), and Langenhoff and Stuckey (2000) showed
that for compartmentalized anaerobic configurations the actual
number of compartments correlates closely with an equal number
of perfectly mixed compartments, and that, as such, plug-flow
conditions were approached. The resulting SCOD and VFA con-
centration gradients at a 4-hour HRT were used as a baseline to
Figure 4 Area distribution of granules from the start, compare reactor performances during and after the hydraulic
middle, and end of the operating period. stress.

572 Water Environment Research, Volume 73, Number 5


Angenent et al.

The 4-hour to 1-hour HRT drop at day 185 increased the VLR stability of the system during changes in flow, as was demonstrated
from 3.5 g COD/Ld to 14.4 g COD/Ld (COD level of the feed during a temporary drop in HRT from 4 hours to 1 hour.
was kept constant). Consequently, the F/M ratio was increased The results of this study indicate that the AMBR is an attractive
from 0.18 to 0.72 g COD/g VSSd (Figure 2c). The SCOD, TCOD, option for treating highly soluble industrial and domestic low-
and MCOD removal efficiencies decreased to 39.3% (eSCOD strength wastewater at low-temperature conditions, because the
384.3 mg/L), 30.3% (eTCOD 441.2 mg/L), and 14.8% (SE AMBR showed high biomass retention and stability during
0.8; n 4), respectively, while the VFA concentration of the changes in flow. Moreover, at a 4-hour HRT it showed substrate
effluent increased to 110 mg/L as acetic acid (Figure 3b). Despite staging with relatively high SCOD and acetic acid concentrations
lower removal efficiencies, the AMBR removed more SCOD in the initial compartment and low SCOD and acetic acid concen-
during than before the hydraulic stress test (123.4 and 58.5 g/d, trations in the final compartment. These conditions were beneficial
respectively). The hydraulic stress did not upset the reactor in in promoting a complete removal of propionic acid in the final
terms of a pH drop in the initial compartment, but reactor perfor- compartments of the reactor. We did not study the dynamic be-
mance (COD removal efficiencies) was deteriorated severely. Sub- havior of the AMBR because of the flow reversals over the
strate staging that was apparent at a 4-hour HRT was lost com- horizontal plane of the reactor and the possible advantages this
pletely at day 185 (Figure 5b). For example, the acetic acid might have on reactor stability. Also, we did not optimize the
concentration in compartment 1 was lower than that in compart- frequency of reversing the flow over the horizontal plane of the
ments 2 to 4. Moreover, propionic acid levels that were lower than reactor. Hence, further studies should include optimization of the
the detection limit in compartments 2 to 4 at a 4-hour HRT, had AMBR operation, while feeding real domestic or low-strength
reached a level as high as 22 mg/L during the hydraulic stress. industrial wastewater.
Clearly, a shift in degradation of nonfat dry milk substrate had
occurred from the initial compartment to the middle and final Acknowledgments
compartments because of the flow increase from 120 to 480 L/d. Credits. This research was made possible by a grant from the
The biomass migration rate during the hydraulic stress increased U.S. Department of Agriculture (Washington, D.C.), through the
from virtually zero to approximately 13.2 g VSS/Ld, and hence Iowa Biotechnology Byproducts Consortia (Ames, Iowa). The
more granules were migrating through the system. Also, the wash authors acknowledge Richard Dague, who initiated anaerobic re-
out of biomass with the effluent increased from 8.7 to 35 g VSS/d search with low-strength wastewater at Iowa State University
(Figure 2b), which decreased the SRT from 50.6 to 12.6 days (Ames) and Warren Straszheim (Civil and Construction Engineer-
(Figure 2c). However, after 1 day of feeding the AMBR at a 1-hour ing Material Analyses Research Laboratory, Iowa State Univer-
HRT, the MLVSS had decreased only by 1 g/L. Because of the sity) for his assistance with image analyses. This paper was pre-
retention of biomass in the system, the reactor performed almost sented in part at WEFTEC 2000, the Water Environment
Federations 73rd Annual Conference and Exposition held in Ana-
similarly the day after the hydraulic stress as compared to the day
heim, California, October 14 18, 2000.
before the hydraulic stress, with SCOD, TCOD, and MCOD re-
Authors. At the time of this work, all authors were in the
moval efficiencies of 69.6% (eSCOD 168.1 mg/L), 49.2%
Department of Civil and Construction Engineering at Iowa State
(eTCOD 280.5), and 32.8%, respectively (Figure 3b). Figure 5c
University. Largus T. Angenent is currently a postdoctoral re-
illustrates that the AMBR showed substrate staging at day 186
search associate in the Department of Civil, Environmental, and
similar to that before the hydraulic stress. Only slightly elevated
Architectural Engineering, University of Colorado at Boulder;
acetic and propionic acid concentrations of approximately 10 mg/L
Gouranga C. Banik is an assistant professor in the construction
in compartments 2 to 4 indicated that the system had been stressed
department at Southern Polytechnic State University (Atlanta,
(Figure 5c). Studies with an alternative compartmentalized anaer-
Georgia); and Shihwu Sung is an assistant professor in the De-
obic system, called the anaerobic baffled reactor, also showed a
partment of Civil and Construction Engineering, Iowa State Uni-
compartmentalized system to be very stable during large changes
versity. Correspondence should be addressed to Largus T. Ange-
in flow (Nachaiyasit and Stuckey, 1997). The anaerobic baffled
nent, Department of Civil, Environmental and Architectural
reactor recovered to its baseline performance shortly after the
Engineering, University of Colorado at Boulder, Campus Box 428,
period of hydraulic stress ended. Hence, anaerobic systems with a
Boulder, CO 80309.
configuration that promotes high biomass retention are required in
Submitted for publication November 16, 2000; revised manu-
off-setting hydraulic stress. This is especially important for low-
script submitted May 14, 2001; accepted for publication August
strength wastewater at low-temperature conditions, because of the
28, 2001
low biomass growth rate. The Deadline to submit Discussions of this paper is February
15, 2002.
Conclusions
A laboratory-scale AMBR, which consisted of four compart- References
ments and was fed with a nonfat dry milk synthetic wastewater at American Public Health Association; American Water Works Association;
a concentration of 600 mg COD/L (BOD5 of 285 mg/L) at a and Water Environment Federation (1995) Standard Methods for the
temperature of 15 C, achieved SCOD, TCOD, and MCOD re- Examination of Water and Wastewater. 19th Ed., Washington, D.C.
Angenent, L.T., and Sung, S. (2001) Development of Anaerobic Migrating
moval efficiencies of 73%, 59%, and 36%, respectively, at a
Blanket Reactor (AMBR), a Novel Anaerobic Treatment System.
4-hour HRT. The reactor removal rate improved over the operating Water Res. (G.B.) 35, 1739.
time, during which the specific methanogenic activity of the gran- Banik, G.C.; Ellis, T.G.; and Dague, R.R. (1997) Structure and Methano-
ules increased slowly. The improvement in removal rate was genic Activity of Granules from an ASBR Treating Dilute Wastewater
possible because of long biomass SRTs in the AMBR resulting at Low Temperatures. Water Sci. Technol. (G.B.), 36, 6/7, 149.
from effective granular retention. Biomass retention also promoted Banik, G.C.; Viraraghavan, T.; and Dague, R.R. (1998) Low Temperature

September/October 2001 573


Angenent et al.

Effects on Anaerobic Microbial Kinetic Parameters. Environ. Tech- the Performance of an Anaerobic Baffled Reactor (ABR). 2. Step and
nol., 19, 5, 503. Transient Hydraulic Shocks at Constant Feed Strength. Water Res.
Barbosa, R.A., and Sant Anna, G.L. (1989) Treatment of Raw Domestic (G.B.), 31, 2747.
Sewage in an UASB Reactor. Water Res. (G.B.), 23, 1483. Perry, R.H.; Green, D.W.; and Maloney, J.O. (1997) Perrys Chemical
Dague, R.R.; Banik, G.C.; and Ellis, G.E. (1998) Anaerobic Sequencing Engineers Handbook. McGraw-Hill, New York.
Batch Reactor Treatment of Dilute Wastewater at Psychrophilic Tem- Rebac, S.; van Lier, J.B.; Janssen, M.G.J., Dekkers, F.; Swinkels, K.T.M.;
peratures. Water Environ. Res., 70, 155. and Lettinga, G. (1997) High-Rate Anaerobic Treatment of Malting
Grady, C.P.L. Jr.; Daigger, G.T.; and Lim, H.C. (1999) Biological Waste- Waste Water in a Pilot-Scale EGSB System Under Psychrophilic
water Treatment. Marcel Dekker, New York. Conditions. J. Chem. Technol. Biotechnol., 68, 135.
Grobicki, A., and Stuckey, D.C. (1992) Hydrodynamic Characteristics of Rinzema, A.; van Lier, J.; and Lettinga, G. (1988) Sodium Inhibition of
the Anaerobic Baffled Reactor. Water Res. (G.B.), 26, 371. Acetoclastic Methanogens in Granular Sludge from a UASB Reactor.
Grotenhuis, J.T.C.; Kissel, J.C.; Plugge, C.M.; Stams, A.J.M.; and Ze- Enzyme Microbiol. Technol., 10, 24.
hnder, A.J.B. (1991) Role of Substrate Concentration in Particle Size Rodrigues, A.C.; Brito, A.G.; and Melo, L.F. (2001) Posttreatment of a
Distribution of Methanogenic Granular Sludge in UASB Reactors. Brewery Wastewater Using a Sequencing Batch Reactor. Water En-
Water Res. (G.B.), 25, 21. viron. Res., 73, 45.
Hulshoff Pol, L.; Euler, H.; Eitner, A.; and Grohganz, D. (1997) GTZ
Sajjad, M.W., and Cleasby, J.L. (1995) Effect of Impeller Geometry and
Sectorial Project: Promotion of Anaerobic Technology for the Treat-
Various Mixing Patterns on Flocculation Kinetics of Kaolin Clay
ment of Municipal and Industrial Sewage and Wastes. Proc. 8th Int.
Using Ferric Salts. Proc. Annu. Conf. Am. Water Works Assoc.,
Conf. Anaerobic Digestion, Sendai, Jpn.; Int. Assoc. Water Qual.,
Denver, Colo.
London.
Sekiguchi, Y.; Kamagata, Y.; Nakamura, K.; Ohashi, A.; and Harada, H.
Kato, M.T. (1994) The Anaerobic Treatment of Low Strength Soluble
(1999) Flourescence In Situ Hybridization Using 16S rRNA-Target-
Wastewaters. Ph. D. thesis, Dep. Environ. Technol., Wageningen
ted Oligonucleotides Reveals Localization of Methanogens and Se-
Agricultural University, The Netherlands.
Langenhoff, A.A.M., and Stuckey, D.C. (2000) Treatment of Dilute Waste- lected Uncultured Bacteria in Mesophilic and Thermophilic Sludge
water Using an Anerobic Baffled Reactor: Effect of Low Tempera- Granules. Appl. Environ. Microbiol., 65, 1280.
ture. Water Res.(G.B.), 34, 3867. Sung, S., and Dague, R.R. (1995) Laboratory Studies on the Anaerobic
Lettinga, G.; de Man, A.; van der Last, A.R.M.; Wiegant, W.; van Knip- Sequencing Batch Reactor. Water Environ. Res., 67, 294.
penberg, K.; Frijns, J.; and van Buuren, J.C.L. (1993) Anaerobic Switzenbaum, M.S., and Grady, C.P.L. Jr. (1986) Anaerobic Treatment of
Treatment of Domestic Sewage and Wastewater. Water Sci. Technol. Domestic Wastewater. J. Water Pollut. Control Fed., 58, 102.
(G.B.), 27, 967. van Lier, J.B.; Groeneveld, N.; and Lettinga, G. (1996) Development of
Lettinga, G.; Rebac, S.; Parshina, S.; Nozhevnikova, A.; van Lier, J.B.; and Thermophilic Methanogenic Sludge in Compartmentalized Upflow
Stams, A.J.M. (1999) High-Rate Anaerobic Treatment of Wastewater Reactors. Biotechnol. Bioeng., 50, 115.
at Low Temperatures. Appl. Environ. Microbiol., 65, 1696. van Lier, J.B.; Rebac, S.; Lens, P.; van Bijnen, F.; Oude Elferink, S.J.W.H.;
Mergaert, K.; Vanderhaegen, B.; and Verstraete, W. (1992) Applicability Stams, A.J.M.; and Lettinga, G. (1997) Anaerobic Treatment of Partly
and Trends of Anaerobic Pre-Treatment of Municipal Wastewater. Acidified Wastewater in a Two-Stage Expanded Granular Sludge Bed
Water Res. (G.B.), 26, 1025. (EGSB) System at 8 C. Water Sci. Technol. (G.B), 36, 6/7, 317.
Monod, J. (1949) The Growth of Bacterial Cultures: a Review. Microbiol., Wirtz, R.A., and Dague, R.R. (1996) Enhancement of Granulation and
3, 371. Start-Up in the Anaerobic Sequencing Batch Reactor. Water Environ.
Nachaiyasit, S., and Stuckey, D.C. (1997) The Effect of Shock Loads on Res., 68, 883.

574 Water Environment Research, Volume 73, Number 5

You might also like